H. Makino , N. Minami , and S. Tasaki arXiv:0902.1282v2 ... · semiclassical periodic-orbit theory...

22
arXiv:0902.1282v2 [nlin.CD] 27 Mar 2009 Statistical properties of spectral fluctuations for a quantum system with infinitely many components H. Makino 1 , N. Minami 2 , and S. Tasaki 3 1 Department of Human and Information Science, Tokai University, 1117 Kitakaname, Hiratsuka-shi, Kanagawa 259-1292, Japan 2 School of Medicine, Keio University, Hiyoshi 4-1-1, Kohoku-ku, Yokohama, Kanagawa 223-8521, Japan 3 Department of Applied Physics, Waseda University, 3-4-1 Okubo, Shinjuku-ku, Tokyo 169-8555, Japan Abstract Extending the idea formulated in Makino et al[Phys.Rev.E 67,066205], that is based on the Berry–Robnik approach [M.V. Berry and M. Robnik, J. Phys. A 17, 2413], we investigate the statistical properties of a two-point spectral correlation for a classically integrable quantum system. The eigenenergy sequence of this system is regarded as a superposition of infinitely many independent components in the semi- classical limit. We derive the level number variance (LNV) in the limit of infinitely many components and discuss its deviations from Poisson statistics. The slope of the limiting LNV is found to be larger than that of Poisson statistics when the individual components have a certain accumulation. This property agrees with the result from the semiclassical periodic-orbit theory that is applied to a system with degenerate torus actions[D. Biswas, M.Azam,and S.V.Lawande, Phys. Rev. A 43, 5694]. 1 typeset using P T P T E X.cls Ver.0.9

Transcript of H. Makino , N. Minami , and S. Tasaki arXiv:0902.1282v2 ... · semiclassical periodic-orbit theory...

  • arX

    iv:0

    902.

    1282

    v2 [

    nlin

    .CD

    ] 2

    7 M

    ar 2

    009

    Statistical properties of spectral fluctuations for a

    quantum system with infinitely many components

    H. Makino1, N. Minami2, and S. Tasaki3

    1Department of Human and Information Science, Tokai University, 1117

    Kitakaname, Hiratsuka-shi, Kanagawa 259-1292, Japan2School of Medicine, Keio University, Hiyoshi 4-1-1, Kohoku-ku,

    Yokohama, Kanagawa 223-8521, Japan3Department of Applied Physics, Waseda University, 3-4-1 Okubo,

    Shinjuku-ku, Tokyo 169-8555, Japan

    Abstract

    Extending the idea formulated in Makino et al[Phys.Rev.E 67,066205], that is

    based on the Berry–Robnik approach [M.V. Berry and M. Robnik, J. Phys. A 17,

    2413], we investigate the statistical properties of a two-point spectral correlation for

    a classically integrable quantum system. The eigenenergy sequence of this system is

    regarded as a superposition of infinitely many independent components in the semi-

    classical limit. We derive the level number variance (LNV) in the limit of infinitely

    many components and discuss its deviations from Poisson statistics. The slope of the

    limiting LNV is found to be larger than that of Poisson statistics when the individual

    components have a certain accumulation. This property agrees with the result from the

    semiclassical periodic-orbit theory that is applied to a system with degenerate torus

    actions[D. Biswas, M.Azam,and S.V.Lawande, Phys. Rev. A 43, 5694].

    1 typeset using PTPTEX.cls 〈Ver.0.9〉

    http://arxiv.org/abs/0902.1282v2

  • §1. Introduction

    Study of energy level statistics has played an important role in elucidating the universal

    properties of quantum systems, which in the semiclassical limit, reflect regular and chaotic

    features of classical dynamics. Berry and Tabor conjectured that the eigenenergy levels of a

    quantum system, whose classical dynamical system is integrable, have the same fluctuation

    property as uncorrelated random numbers from a Poisson process, and thus are character-

    ized by Poisson statistics.1) This conjecture is in contrast with the conjecture of Bohigas,

    Giannoni, and Schmit, which assert that GOE or GUE statistics based on random matrix

    theory (RMT) are applicable to the fluctuation property of energy levels of a quantum sys-

    tem whose classical dynamical system is fully chaotic.2) These contrasting conjectures have

    been examined using various statistical observables, e.g., the nearest-neighbor level-spacing

    distribution (NNLSD) and the level number variance (LNV).3)

    The NNLSD P (S) is the observable most commonly used to study short-range fluctua-

    tions in a spectrum. For eigenenergy levels on the unfolded scale,4) this observable is defined

    as the probability density of finding a distance S between adjacent levels. In Poisson statis-

    tics, it is characterized by the exponential distribution P (S) = e−S; on the other hand, in

    GOE/GUE statistics, it approximates the Wigner distribution very well.3)

    The LNV Σ2(L) is the observable commonly used to study correlations between pairs of

    levels, which characterizes long-range fluctuations in a spectrum. It is defined as the average

    variance of the number of levels in an energy interval containing an average of L levels. On

    the unfolded scale, this interval is equivalent to an interval of length L, and the LNV is thus

    defined as

    Σ2(L) = 〈(N (ǫ+ L)−N (ǫ)− L)2〉, (1.1)

    where N (ǫ) is the number of eigenenergy levels below ǫ, and the brackets 〈· · · 〉 represent

    the average over the value of ǫ. In Poisson statistics, the LNV is equal to the number it-

    self [i.e., Σ2Poisson

    (L) = L]; on the other hand, in RMT, in which β = 1 and 2 correspond to

    the GOE and GUE level statistics, respectively, the LNV increases logarithmically with L[

    i.e., Σ2β(L) ∼ (2/βπ2) ln(2πL)

    ]

    .3) According to Berry’s semiclassical theory,5) the LNV of

    a quantum system with f degrees-of-freedom should display these universal statistics in the

    region L

  • to be characterized by Poisson statistics is now widely accepted as a universal property of

    generic integrable quantum systems in the semiclassical limit.

    One possible mechanism producing Poisson statistics has been proposed by Makino et

    al ,19) based on the Berry-Robnik approach.20) We briefly review the outline below: For an in-

    tegrable system, individual orbits are confined in each inherent torus whose surface is defined

    by holding its action variable constant, and the whole region of the phase space is densely

    covered with invariant tori. In other words, the phase space consists of infinitely many re-

    gions, which have infinitesimal volumes in Liouville measure. Because of the suppression of

    quantum tunneling in the semiclassical limit, the Wigner functions of quantal eigenstates are

    expected to be localized in the phase space region explored by a typical trajectory,21), 22) and

    to form independent components. For a classically integrable quantum system, the Wigner

    function localizes on the infinitesimal region in ~ → 0 and tends to a δ function on a torus

    that is designated by quantum numbers.23) Then, the eigenenergy levels can be represented

    as a statistically independent superposition of infinitely many components, each of which

    gives an infinitesimal contribution to the level statistics. Therefore, if the individual spectral

    components are sparse enough, one would expect Poisson statistics to apply here, as a result

    of the law of small numbers.24)

    The statistical independence of spectral components is assumed to be justified by the

    principle of uniform semiclassical condensation of eigenstates in the phase space and by the

    lack of their mutual overlap, and thus, can only be expected in the semiclassical limit.22), 25)

    This mechanism was introduced as a basis for the Berry-Robnik approach to investigate

    the relation between the phase space geometry and the formation of energy level sequence

    in a generic mixed quantum system, whose classical dynamics is partly regular and partly

    chaotic.20) The validity of the Berry-Robnik approach is confirmed by numerical computa-

    tions for the mixed quantum systems in the extremely deep semiclassical region,26), 27) which

    is called the Berry-Robnik regime.28)

    Based on this view, Makino and Tasaki investigated the short-range spectral statistics

    of classically integrable systems.19) They derived the cumulative function of NNLSD, i.e.,

    M(S) =∫ S

    0P (x)dx, is derived in the limit of infinitely many components, which is charac-

    terized by a single monotonically increasing function µ̄(0;S) of the nearest level spacing S

    as

    Mµ̄(S) = 1− [1− µ̄(0;S)] exp

    (

    ∫ S

    0

    [1− µ̄(0; x)] dx

    )

    . (1.2)

    The function µ̄(0;S) classifies Mµ̄(S) into three cases: Case 1, Poisson distribution Mµ̄(S) =

    1 − e−S for ∀S ≥ 0 if µ̄(0; +∞) = 0; Case 2, asymptotic Poisson distribution, which con-

    verges to the Poisson distribution for S → +∞, but possibly not for small spacings S if

    3

  • 0 < µ̄(0; +∞) < 1; and Case 3, sub-Poisson distribution, which converges to 1 for S → +∞

    more slowly than does the Poisson distribution, if µ̄(0; +∞) = 1. Therefore, the Berry-

    Robnik approach when applied to classically integrable quantum systems, admits deviations

    from Poisson statistics.

    Cases 2 and 3 are possible when the eigenenergy levels of individual components show

    a singular NNLSD result from strong accumulation19) (see also Section 3 of the present

    paper), which is expected when the classical dynamical systems have a spatial symme-

    try1), 11), 14), 17), 18), 29), 30), 31), 32) or a time-reversal symmetry.33), 34), 35) One possible example

    is a rectangular billiard with a rational ratio of squared sides.1), 11), 18), 31), 32) These results

    suggest the existence of a new statistical law different from Poisson statistics in the strongly-

    degenerate quantum systems, and raises the question as to whether similar behaviors appear

    in the two-point spectral correlation.

    Deviation from Poisson statistics, resulting from the symmetry, is also exhibited within

    the framework of the periodic-orbit theory.36) Based on the semiclassical theory of Berry

    and Tabor,7) Biswas, Azam, and Lawande investigated the LNV for classically integrable

    systems with degeneracy in orbit actions.32) They showed that the slope gav(L) of the LNV

    Σ2(L) = gav(L)L is described by the average degeneracy of actions of the periodic orbits,

    which for the degenerated quantum systems, is greater than 1, and it is the slope of the

    variance of Poisson statistics. This result leads to the possibility that the degeneracy of ac-

    tions, induced by the symmetry, could be an essential factor for the singularity of individual

    spectral components that yields Cases 2 and 3. This possibility is confirmed by examining

    the properties of LNV for Cases 2 and 3.

    In this study, extending the theory of Makino and Tasaki,19) we investigate the LNV

    Σ2(L) of quantum systems whose energy levels consist of infinitely many independent com-

    ponents, and show that the non-Poisson limits (Cases 2 and 3) are possibly observed also in

    the long-range spectral fluctuations.

    Based on the Berry-Robnik approach, the overall LNV is derived as follows: We con-

    sider a system whose classical phase space is decomposed into N disjoint regions that give

    the distinct spectral components. The Liouville measures of these regions are denoted by

    ρn(n = 1, 2, 3, · · · , N), which satisfy∑N

    n=1 ρn = 1. Let E(k;L), k = 0, 1, 2, · · · be the distri-

    bution function, which denotes the probability to find k levels in an interval (0, L).3), 37), 38)

    The LNV Σ2(L) is expressed by E(k;L) as

    Σ2(L) =+∞∑

    k=0

    (k − L)2E(k;L). (1.3)

    4

  • Let P (k;S), k = 0, 1, 2, · · · , be the level-spacing distribution, which denotes the probability

    density to find k levels in an interval of length S beginning at an arbitrary level ǫi and ending

    at the level ǫi+k+1. P (k;S) is related to E(k;L) as

    P (k;L) =∂2

    ∂L2

    k∑

    j=0

    (k − j + 1)E(j;L), (1.4)

    and

    E(k;L) =

    ∫ +∞

    L

    dx

    ∫ +∞

    x

    [P (k;S)− 2P (k − 1;S) + P (k − 2;S)] dS, (1.5)

    where P (j < 0;S) = 0.3), 37), 38) Eqs.(1.4) and (1.5) are known as the formulae in the the-

    ory of point process, and are derived as corollaries of the Palm-Khinchin formula.39) The

    NNLSD, which was expressed by P (S) at the beginning of this section, is the special case

    with k = 0.

    When the entire sequence of energy levels is a product of statistically independent su-

    perposition of N subsequences, E(k;L) is decomposed into those of subsequences, en(k;L)

    as3), 40)

    EN(k;L) =∑

    PNn=1 kn=k

    N∏

    n=1

    en(kn; ρnL), (1.6)

    where en satisfies the normalization conditions

    +∞∑

    k=0

    en(k; ρnL) = 1 (1.7)

    and+∞∑

    k=0

    ken(k; ρnL) = ρnL. (1.8)

    In terms of normalized level-spacing distribution pn(k;S) of the subsequence, en(k;L) is

    specified as

    en(k;L) = ρn

    ∫ +∞

    L

    dx

    ∫ +∞

    x

    [pn(k;S)− 2pn(k − 1;S) + pn(k − 2;S)] dS, (1.9)

    where pn(j < 0;S) = 0, and pn(k;S) satisfies the normalizations

    ∫ +∞

    0

    pn(k;S)dS = 1 (1.10)

    and∫ +∞

    0

    Spn(k;S)dS =k + 1

    ρn. (1.11)

    5

  • Note that in general, the individual components are not always unfolded automatically even

    when the overall spectrum is unfolded. However, in a sufficiently small interval [ǫ, ǫ +∆ǫ],

    each spectral component obeys the same scaling law (see Appendix of Makino et al19)) and

    thus is unfolded automatically by an overall unfolding procedure. In the Berry-Robnik ap-

    proach, Eq.(1.6) relates the level statistics in the semiclassical limit with the phase space

    geometry.

    In most general cases, the individual components might have degeneracy of levels that

    lead to singular level spacing distributions. In such a case, it is convenient to use its cumu-

    lative distribution function µn(k;S),

    µn(k;S) =

    ∫ S

    0

    pn(k; x)dx. (1.12)

    This function satisfies

    µn(k − 1;S) ≥ µn(k;S) for all k ≥ 1 and for all S ≥ 0, (1.13)

    and the normalization condition

    +∞∑

    k=0

    [µn(k − 1;S)− µn(k;S)] = 0, (1.14)

    with µn(j = −1;S) = 0, since fn(k;S),

    fn(k;S) =

    {

    µn(k − 1;S)− µn(k;S), k ≥ 1

    1− µn(0;S), k = 0

    denotes the probability to find k levels in an interval of length S beginning at an arbitrary

    level, and obviously satisfies, from the definition of this function, the conditions fn(k;S) ≥ 0

    and∑+∞

    k=0 fn(k;S) = 1.

    In addition to Eq.(1.6), we introduce the following two assumptions that were introduced

    in Ref.:19)

    Assumption (i). The statistical weights of independent regions vanish uniformly in the

    limit of infinitely many regions:

    maxn

    ρn → 0 as N → +∞. (1.15)

    Assumption (ii). For k = 0, 1, 2, · · · , the weighted mean of the cumulative distribution of

    energy spacing, namely,

    µ(k; x) =N∑

    n=1

    ρnµn(k; x) (1.16)

    6

  • converges as N → +∞ to µ̄(k; x),

    limN→+∞

    µ(k; x) = µ̄(k; x), (1.17)

    where the convergence is uniform on each closed interval: 0 ≤ x ≤ S. In the Berry-Robnik

    approach, the statistical weights of individual components are the phase volumes (Liouville

    measures) of the corresponding invariant regions.

    Under the Assumptions (i) and (ii), Eqs.(1.3) and (1.6), we obtain the overall LNV in

    the limit N → +∞:

    Σ2µ̄(L) = L+ 2

    ∫ L

    0

    +∞∑

    k=0

    µ̄(k;S)dS. (1.18)

    When µ̄(k;S) = 0 for all k, the LNV of the whole energy sequence reduces to Σ2Poisson

    (L) =

    L. This condition is expected when the individual components are sufficiently sparse. In

    general, we expect µ̄(k;S) > 0, which corresponds to a certain accumulation of the levels of

    individual components. In this case, the LNV of the whole energy sequence deviates from

    Σ2Poisson

    (L) in such a way that the slope of the variance is greater than 1.

    The present paper is organized as follows. The limiting LNV (1.18) is derived from

    Eqs.(1.3) and (1.6), and Assumptions (i) and (ii) in Section 2. In Section 3, the property of

    the limiting LNV is analyzed for the Cases 1–3, wherein the deviations from Poisson statistics

    are clearly observed in Cases 2 and 3. We present a numerical analysis for the rectangular

    billiard that shows deviations from Poisson statistics in Section 4. In the concluding section,

    we discuss some relationships between our results and other related works.

    §2. limiting level number variance

    In this section, by using Eqs.(1.6) and (1.3), and Assumptions (i) and (ii) introduced in

    Section1, we derive the limiting LNV,

    Σ2µ̄(L) = L+ 2

    ∫ L

    0

    +∞∑

    k=0

    µ̄(k;S)dS, (2.1)

    in the limit of infinitely many components N → +∞.

    First, we rewrite Eq.(1.3) in terms of the function en, and decompose it into the LNV σ2n

    of individual components:

    Σ2N(L) =

    +∞∑

    k=0

    (k − L)2∑

    PNn=1 kn=k

    N∏

    n=1

    en(kn; ρnL) (2.2)

    =N∑

    n=1

    +∞∑

    k=0

    (k − ρnL)2en(k; ρnL) ≡

    N∑

    n=1

    σ2n(ρnL), (2.3)

    7

  • where we have used the properties∑+∞

    k=0EN (k;L) = 1,∑+∞

    k=0 kEN(k;L) = L, and

    +∞∑

    k=0

    k2EN (k;L) =N∑

    n=1

    +∞∑

    kn=0

    k2nen(kn; ρnL) + L2 −

    N∑

    n=1

    (ρnL)2, (2.4)

    which follow from Eqs.(1.7) and (1.8), and the relationships k =∑N

    n=1 kn and∑N

    n=1 ρn = 1

    (see also Appendix A). The functions en(k; ρnL) in the above equation are rewritten in terms

    of the cumulative level-spacing distribution functions µn(k;S) of individual components.

    en(k; ρnL) =

    −ρn∫ +∞

    LdS [µn(k;S)− 2µn(k − 1;S) + µn(k − 2;S)] , k ≥ 2

    −ρn∫ +∞

    LdS [µn(1;S)− 2µn(0;S) + 1] , k = 1

    ρn∫ +∞

    LdS [1− µn(0;S)] , k = 0

    (2.5)

    =

    ρn∫ L

    0dS [µn(k;S)− 2µn(k − 1;S) + µn(k − 2;S)] , k ≥ 2

    ρn∫ L

    0dS [µn(1;S)− 2µn(0;S) + 1] , k = 1

    1− ρn∫ L

    0dS [1− µn(1;S)] , k = 0

    (2.6)

    where Eq.(2.6) follows from Eq.(1.11), integration by parts and the following limits which

    result from the existence of the average:

    limS→+∞

    S [µn(k;S)− 2µ(k − 1;S) + µn(k − 2;S)] = 0, k ≥ 2

    limS→+∞

    S [µn(1;S)− 2µ(0;S) + 1] = 0, k = 1 (2.7)

    limS→+∞

    S [1− µn(0;S)] = 0, k = 0

    Then, σ2n(ρnL) is described in terms of µn(k;S) as

    σ2n(ρnL) =+∞∑

    k=0

    k2en(k; ρnL)− ρ2nL

    2

    = ρn

    ∫ L

    0

    dS

    +∞∑

    k=1

    (2k + 1) [µn(k − 1;S)− µn(k;S)] + ρn

    ∫ L

    0

    [1− µn(0;S)] dS − ρ2nL

    2

    = ρnL+ 2

    ∫ L

    0

    dS

    +∞∑

    k=0

    ρnµn(k;S)− ρ2nL

    2, (2.8)

    where we have used Eq.(1.14) and the relation

    +∞∑

    k=0

    [(k − 1)µn(k − 1;S)− kµn(k;S)] = 0 (2.9)

    with µn(−1;S) = 0. Therefore, in the limit N → +∞, we have the convergence

    Σ2µ̄(L) = limN→+∞

    N∑

    n=1

    σ2n(ρnL) = L+ 2

    ∫ L

    0

    dS+∞∑

    k=0

    µ̄(k;S), (2.10)

    8

  • where the limit (2.10) follows from Assumption (ii) and the following property results from

    Assumption (i):N∑

    n=1

    ρ2n ≤ maxn

    ρn

    N∑

    n=1

    ρn = maxn

    ρn → 0. (2.11)

    We note the spectral rigidity ∆3(L), which was introduced by Dyson and Mehta.41)

    In combination with the LNV, this quantity has played a major role in the study of the

    long-range spectral statistics. According to Pandey42) and Eq.(2.10), the spectral rigidity is

    described in terms of µ̄(k;S) as

    ∆3,µ̄(L) =2

    L4

    ∫ L

    0

    dx(L3 − 2L2x+ x3)Σ2µ̄(x)

    =L

    15+

    4

    L4

    ∫ L

    0

    dx(L3 − 2L2x+ x3)

    ∫ x

    0

    dS+∞∑

    k=0

    µ̄(k;S). (2.12)

    §3. Properties of limiting level number variance

    We analyze the slope of the limiting LNV

    g(L) =1

    LΣ2µ̄(L) = 1 +

    2

    L

    ∫ L

    0

    dS

    +∞∑

    k=0

    µ̄(k;S), (3.1)

    which has the following convergence:

    limL→+∞

    g(L) = 1 + 2

    +∞∑

    k=0

    µ̄(k; +∞) ≥ 1 + 2µ̄(0; +∞). (3.2)

    Since µn(k;S) is monotonically increasing for S ≥ 0, monotonically decreasing for k =

    0, 1, 2, · · · , and 0 ≤ µn(k;S) ≤ 1, µ̄(k;S) has the same properties and is bounded by

    µ̄(0; +∞) as

    0 ≤ µ̄(k;S) ≤ µ̄(0;S) ≤ µ̄(0; +∞) ≤ 1, (3.3)

    where µ̄(0; +∞) classifies the cumulative NNLSD (1.2) into three cases: Case 1, the Poisson

    distribution if µ̄(0; +∞) = 0; Case 2, the asymptotic Poisson distribution if 0 < µ̄(0; +∞) <

    1; and Case 3, the sub-Poisson distribution if µ̄(0; +∞) = 1.

    Then, the property of the limiting LNV is evaluated for Cases 1—3 as follows:

    Case 1, µ̄(0; +∞) = 0: The limiting LNV Σ2µ̄(L) agrees with the LNV of Poisson statis-

    tics: Σ2Poisson

    (L) = L. Note that this condition is equivalent to∑+∞

    k=0 µ̄(k;S) = 0 since µ̄(k;S)

    is monotonically increasing for S and decreasing for k.

    Case 2, 0 < µ̄(0; +∞) < 1: Σ2µ̄(L) possibly deviates from Σ2Poisson

    (L) in such a way that

    the slope of the limiting LNV is 1 at L = 0, increases monotonically with L, and approaches

    9

  • a number 1 + 2µ̄(0; +∞) or more as L → +∞.

    Case 3, µ̄(0; +∞) = 1: Σ2µ̄(L) possibly deviates from Σ2Poisson

    (L) in such a way that the

    slope of the limiting LNV is 1 at L = 0, increases monotonically with L, and approaches a

    number 3 or more as L → +∞.

    One has Case 1 if the NNLSD of individual components are derived from the scaled

    distribution functions ϕn(0;S) as

    µn(0;S) = ρn

    ∫ S

    0

    ϕn(0; ρnx)dx, (3.4)

    where ϕn(0; ρnS) = pn(0;S)/ρn and satisfy∫ +∞

    0

    ϕn(0; x)dx = 1,

    ∫ +∞

    0

    xϕn(0; x)dx = 1, (3.5)

    and are uniformly bounded by a positive constant D : |ϕn(0;S)| ≤ D ( 1 ≤ n ≤ N ). Indeed,

    the following holds:

    |µ(0;S)| ≤N∑

    n=1

    ρ2n

    ∫ S

    0

    |ϕn(0; ρnx)| dx ≤ DSN∑

    n=1

    ρ2n ≤ DSmaxn

    ρn

    N∑

    n=1

    ρn → 0 ≡ µ̄(0;S).

    (3.6)

    Such a bounded condition is possible when the individual spectral components are sparse

    enough.

    In general, one may expect Cases 2 or 3 with µ̄(0;S) > 0, which corresponds to strong

    accumulation of energy levels, leading to a singular NNLSD of the individual components.

    Such accumulation is expected to arise from the symmetry of the system. In the next section,

    we will analyze the LNV of the rectangular billiard systems which is known to deviate from

    Poisson statistics.

    §4. Rectangular billiard system

    We present our results on the various statistical measures discussed in the previous section

    for a rectangular billiard system whose spectral statistics has been precisely analyzed in a

    number of works.1), 5), 11), 15), 18), 19), 31), 32) The eigenenergy levels of this system are given by

    ǫn,m = n2 + αm2, (4.1)

    where n and m are positive integers, and α is denoted by the lengths of two sides a and b

    as α = a2/b2. The unfolding transformation {ǫn,m} → {ǭn,m} is carried out by using the

    leading Weyl term of the integrated density of states, N (ǫ), as

    ǭn,m = N (ǫn,m) =π

    4√

    αǫn,m. (4.2)

    10

  • Berry and Tabor observed that the NNLSD of this system possibly deviates from Poisson

    statistics when α is rational.1) In this paper, we study irrational cases in addition to a

    rational case(α = 1) that are described by a finite continued fraction of the golden mean

    (√

    5 + 1)/2,

    α = 1 +1

    1+

    1

    1+· · ·

    1

    1+

    1

    1 + δ= [1; 1, 1, · · · , 1, 1 + δ] , (4.3)

    with an irrational truncation parameter δ ∈ [0, 1).

    Fig.1 shows the plots of the LNV Σ2(L) for α corresponding to the (a) 25th, (b) 8th,

    and (c) 4th approximations of the golden mean, and (d) α = 1. Our analysis is valid in

    the region L

  • µ̃(S) in S → +∞ approaches a number µ̃(+∞) such that 0 < µ̃(+∞) < 1 [plots (b) – (d)

    in Fig.4]. Therefore, plots (b)–(d) correspond to Case 2.

    For the rectangular billiard in the finite energy region, we have not yet succeeded in

    observing Case 3. This case is expected to arise in a square billiard (α = 1) in the high

    energy limit where a stronger accumulation of levels is generated. Based on the number-

    theoretical result of Landau (1908),43) Connors and Keating have proved that the eigenenergy

    levels ǫn,m of square billiard show a logarithmic increase in the mean degeneracy of levels as

    ǫ → +∞, which is described as

    1−M(+0) ≃4

    π

    C2√

    ln ǫ→ 0, (4.5)

    where C2 converges to give C2 ≃ 0.764.11) The above limit corresponds to a delta function

    of NNLSD, P (S) = δ(S), and is consistent with µ̄(0;S = +0) = limǫ→+∞M(+0) = 1, which

    indicates an extremely slow approach to Case 3 in ǫ → +∞.

    Fig.5 shows the cumulative NNLSD M(S) for α = 1. This function is not smooth at the

    level spacings separated by a step π/4,1) which correspond to accumulation of levels. Note

    thatM(+0) > 0 due to the degeneracy at S = 0. Since µ̃(+0) = M(+0) > 0, this degeneracy

    at S = 0 is identified also in Fig.4. As the eigenenergy levels become higher, M(+0) increases

    monotonically and approaches 1. In the limit ǫ → +∞ where M(+0) = µ̄(0, S = +0) = 1,

    all steps except the step at S = 0 are suppressed since M(S) is monotonically increasing for

    S > 0.

    Fig.6 shows 1 − M(+0) vs 4C2/π√

    ln ǫ for various energy ranges. Although we are

    not yet far enough in the high energy region where 1 − M(+0) = 1 − µ̃(+0)

  • many independent components, and each of which gives an infinitesimal contribution to the

    level statistics. Moreover, by developing the approach of Makino et al19) into the statistics

    of higher order (long-range) spectral fluctuations, the LNV of systems consisting of infinitely

    many spectral components are obtained. The LNV is characterized by the monotonically

    increasing functions µ̄(k;S), k = 0, 1, 2, · · · , of the level-spacing S, where the lowest order

    term µ̄(0;S) is associated with the NNLSD. The property of the LNV is classified as follows:

    Case 1, µ̄(0; +∞) = 0 where the cumulative NNLSD is the Poisson distribution, the LNV is

    the Poissonian Σ2Poisson

    (L) = L; Case 2, 0 < µ̄(0; +∞) < 1 where the cumulative NNLSD is

    the asymptotic Poisson distribution, the LNV deviates from Σ2Poisson

    (L) in such a way that

    the slope is greater than 1 and approaches a number ≥ 1 + 2µ̄(0; +∞) as L → +∞; Case

    3, µ̄(0; +∞) = 1 where the cumulative NNLSD is the sub-Poisson distribution, the LNV

    deviates from Σ2Poisson

    (L) in such a way that the slope is greater than 1 and approaches a

    number ≥ 3 as L → +∞. Therefore, we have shown that deviations from Poisson statistics

    (Cases 2 and 3) are possibly observed, not only in the property of NNLSD characterizing

    short-range spectral fluctuation as shown in the work of Makino et al ,19) but also in the

    property of the LNV characterizing the fluctuations of all ranges.

    Note that Cases 2 and 3 may arise when there is a strong accumulation of levels, which is

    characterized by the singular NNLSD of individual components. Such accumulation would

    be expected when there is a spatial symmetry or time-reversal symmetry. One example is

    a rectangular billiard shown in Section 4 where the result shows Case 2 in addition to Case

    1, and an extremely slow approach to Case 3 . Similar results due to a spatial symmetry

    are also shown in the equilateral-triangular billiard,30), 44) torus billiard,29), 18) and integrable

    Morse oscillator.31) Another example is a certain type of systems with the time-reversal

    symmetry studied by Shnirelman,33) Chirikov and Shepelyansky,34) and Frahm and Shep-

    elyansky.35) In this system, the strong accumulation of energy levels, resulting from the time

    reversibility, is reflected in the NNLSD as a sharp Shnirelman peak at small level-spacings.

    Rigorous results confirming the Berry–Tabor conjecture (Case 1) are also reported for a

    certain classically integrable system. In the work of Marklof13) and Eskin et al ,14) eigenvalue

    problems are reformulated as lattice point problems, and it is exactly proved under explicit

    diophantine conditions that a two-point spectral correlation exhibits Poisson statistics.

    The periodic-orbit theory, applied for an integrable system with degeneracy in orbit ac-

    tions, gives a similar result. In the work of Biswas et al ,32) it was shown that the slope gav(L)

    of the LNV Σ2(L) = gav(L)L represents the average degeneracy of actions of periodic orbits,

    which is greater than 1 for the rectangular billiard with a rational ratio α of squared sides.

    Since this property is qualitatively consistent with the property of the limiting LNV (2.10)

    13

  • in Cases 2 and 3, one might expect to have the relation:

    gav(L) = 1 +2

    L

    ∫ L

    0

    +∞∑

    k=0

    µ̄(k;S)dS, (5.1)

    which enables us to discuss the properties of the overall NNLSD from the periodic-orbit

    theory. From Eqs.(5.1) and (3.3), we expect that the Case 1 corresponds to the non-

    degeneracy of actions: gav(L) = 1, while Cases 2 and 3 correspond to the degeneracy of

    actions:gav(L) > 1. Therefore, we have confirmed that Cases 2 and 3 are closely related to

    the degeneracy in orbit actions. At the same time, we have only superficially examined the

    possibility that the degeneracy of actions, induced by the symmetry, could be an essential

    factor for the singularity of individual spectral components that yields Cases 2 and 3. This

    part should be investigated in detail in a future work.

    There is another approach similar to the one presented in this paper, in which the LNV

    is derived in the limit of infinitely many components.45) The LNV is described by the Dyson

    two-level cluster function Y2 as

    Σ2(L) = L− 2

    ∫ L

    0

    (L− S)Y2(S)dS. (5.2)

    For spectral superposition, Y2 is described in terms of the cluster function of individual

    components y2,n as in Pandey’s work42)

    Y2(S) =N∑

    n=1

    ρ2ny2,n(ρnS). (5.3)

    Then, for the overall LNV, one has the convergence

    limN→+∞

    Σ2(L) = L+ 2

    ∫ L

    0

    c̄(S)dS (5.4)

    with

    c̄(S) = − limN→+∞

    N∑

    n=0

    ρn

    ∫ ρnS

    0

    y2,n(x)dx, (5.5)

    where c̄(S) = 0 corresponds to Poisson statistics and c̄(S) 6= 0 indicates deviations from

    Poisson statistics. Since y2,n is associated with the level-spacing distribution of individual

    components as y2,n(x) = 1−∑+∞

    k=0 pn(k; x), c̄(S) is rewritten as

    c̄(S) = limN→+∞

    N∑

    n=0

    ρn

    ∫ ρnS

    0

    +∞∑

    k=0

    pn(k; x)dx. (5.6)

    14

  • When pn(k;S), k = 0, 1, 2, · · · , are assumed to satisfy the commutation relation∫ ρnx

    0

    ∑+∞k=0 pn(k;S)dS =

    ∑+∞k=0

    ∫ ρnx

    0pn(k;S)dS, c̄(S) is described as c̄(S) =

    ∑+∞k=0 µ̄(k;S) and limit (5

    .4) is consistent

    with the limit (2.10).

    Acknowledgment

    H.M. is grateful to Prof. M. Robnik, Prof. A. Shudo, and Prof. Y. Aizawa for helpful sugges-

    tions. This work is partially supported by KAKENHI(No.18740241 to H.M., No.17540100 to

    N.M., No.17540365 to S.T.), and by a Grant-in-Aid for the “Academic Frontier” Project at

    Waseda University from the Ministry of Education, Culture, Sports, Science and Technology

    of Japan.

    Appendix A

    Derivation of Eqs.(2.3) and (2.4)

    We briefly show the detailed process to derive Eqs.(2.3) and (2.4) from Eqs.(1.4), (1.7),

    and (1.8). First, we rewrite Eq.(1.3) as

    +∞∑

    k=0

    (k − L)2EN (k;L) =+∞∑

    k=0

    k2EN(k;L)− L2, (A.1)

    where we have used the normalization conditions∑+∞

    k=0EN(k;L) = 1 and∑+∞

    k=0 kEN (k;L) =

    L, which follow from Eqs.(1.7) and (1.8), and the relations k =∑N

    n=1 kn and∑N

    n=1 ρn = 1

    as+∞∑

    k=0

    EN(k;L) =N∏

    n=1

    +∞∑

    kn=0

    en(kn; ρnL) = 1, (A.2)

    +∞∑

    k=0

    kEN(k;L) =

    N∑

    n=1

    +∞∑

    kn=0

    knen(kn; ρnL)

    N∏

    i=1,i 6=n

    +∞∑

    ki=0

    ei(ki; ρiL)

    =N∑

    n=1

    ρnL · 1N−1 = L. (A.3)

    By using Eqs.(A.2) and (A.3),∑+∞

    k=0 k2EN(k;L) in the right hand side of Eq.(A.1) is calcu-

    lated in a similar manner:

    +∞∑

    k=0

    k2EN (k;L)

    =

    N∑

    n=1

    +∞∑

    PNi=1,i6=n ki=0

    (

    N∏

    i=1,i 6=n

    ei(ki; ρiL)

    )

    +∞∑

    kn=0

    k2nen(kn; ρnL)

    15

  • +

    N∑

    n 6=i

    +∞∑

    kn=0

    +∞∑

    ki=0

    knen(kn; ρnL)kiei(ki; ρiL)

    =N∑

    n=1

    (

    +∞∑

    kn=0

    k2nen(kn; ρnL)

    )(

    +∞∑

    ki=0

    ei 6=n(ki; ρiL)

    )N−1

    (A.4)

    +

    (

    N∑

    n=1

    +∞∑

    kn=0

    knen(kn; ρnL)

    )2

    −N∑

    n=1

    (

    +∞∑

    kn=0

    knen(kn; ρnL)

    )2

    =N∑

    n=1

    +∞∑

    kn=0

    k2nen(kn; ρnL) + L2 −

    N∑

    n=1

    (ρnL)2. (A.5)

    Since the off-diagonal terms n 6= i vanish in the second equality, we have

    +∞∑

    k=0

    k2EN(k;L)− L2 =

    N∑

    n=1

    (

    +∞∑

    kn=0

    k2nen(kn; ρnL)− (ρnL)2

    )

    =

    N∑

    n=1

    +∞∑

    kn=0

    (kn − ρnL)2en(kn; ρnL). (A.6)

    Therefore, the overall LNV of the system with N independent components is described as

    Σ2N(L) =N∑

    n=1

    σ2n(ρnL), (A.7)

    where σ2n is the LNV of the spectral component:

    σ2n(ρnL) =+∞∑

    k=0

    (k − ρnL)2en(k; ρnL). (A.8)

    References

    1) M.V. Berry and M. Tabor, Proc. R. Soc. Lond. A 356, 375 (1977).

    2) O. Bohigas, M.J. Giannoni, and C. Schmit, Phys. Rev. Lett. 52, 1 (1984).

    3) M.L. Mehta,Random Matrices(2nd ed., San Diego, CA: Academic Press, 1991)

    4) O. Bohigas, Random matrices and chaotic dynamics(LesHouches 1989 Session LII,

    chaos and quantum physics, North-Holland).

    5) M. V. Berry, Proc. Roy. Soc. Lond. A 400, 229 (1985).

    6) J. H. Hannay and A. M. Ozorio de Almedia, J. Phys. A 17, 3429 (1984).

    7) M.V. Berry and M. Tabor, Proc. R. Sco. 349, 101 (1976); M.V. Berry and M. Ta-

    bor, J. Phys. A 10, 371 (1977).

    8) S. A. Molchanov, Commun. Math. Phys. 78, 429 (1981); S. A. Molchanov, Math. USSR

    Izvestija 12, 69 (1978).

    16

  • 9) P. M. Bleher, J. Stat. Phys. 61, 869 (1990); P. M. Bleher, J. Stat. Phys. 63, 261 (1991).

    10) P.Sarnak, Values at integers of binary quadratic forms, Harmonic Analysis and Num-

    ber Theory (Montreal, PQ, 1996), CMS Conf.Proc.21(Amer. Math. Soc., Providence,

    RI.), 181 (1997).

    11) R. D. Connors and J. P. Keating, J. Phys. A 30, 1817 (1997).

    12) J. Marklof, Comm. Math. Phys. 199, 169 (1998).

    13) J. Marklof, Progr. Math.202, Birkhauser,Basel, 421 (2001); DUKE Math.

    J. 115, 409 (2002); Ann. of Math. 158, 419 (2003).

    14) A. Eskin, G.A. Margulis and S. Mozes, Quadratic forms of signature (2,2) and eigen-

    value spacings on rectangular 2-tori, to appear in Ann. Math.

    15) G. Casati, B. V. Chirikov and I. Guarneri, Phys. Rev. Lett. 54, 1350 (1985).

    16) M. Feingold, Phys. Rev. Lett.55, 2626 (1985)

    17) T. H. Seligman and J. J. M. Verbaarschot, Phys. Rev. Lett. 56, 2767 (1986).

    18) M. Robnik and G. Veble, J. Phys. A 31, 4669 (1998).

    19) H. Makino and S. Tasaki, Phys. Rev. E 67, 066205 (2003).

    20) M.V. Berry and M. Robnik, J. Phys. A 17,2413 (1984).

    21) M. V. Berry,J. Phys. A 10, 2083 (1977).

    22) M. Robnik, Nonlinear Phenomena in Complex Systems 1, n1, 1 (1998).

    23) M. V. Berry, Phil. Trans. R. Soc. A 287, 237 (1977).

    24) W. Feller, An introduction to probability theory and its applications (2nd ed., John

    Wiley & Sons, Inc., New York, 1957)

    25) B. Li and M. Robnik, J. Phys. A 27, 5509 (1994).

    26) T. Prosen and M. Robnik, J. Phys. A 26, 2371 (1993); T. Prosen and M. Robnik,

    J. Phys. A 27, 8059 (1994).

    27) T. Prosen, Physica D 91, 244 (1996).

    28) M. Robnik and T. Prosen, J. Phys. A 30, 8787 (1997).

    29) P. J. Richens and M. V. Berry, Physica (Amsterdam) 2D, 495 (1981).

    30) M. V. Berry and M. Wilkinson, Proc. R. Soc. London A 392, 15 (1984).

    31) A. Shudo, Prog. Theor. Phys. Suppl.98, 173 (1989).

    32) D. Biswas, M. Azam, and S.V. Lawande, Phys. Rev. A 43, 5694 (1991); D. Biswas,

    M. Azam, and S.V. Lawande, J. Phys. A 24, 1825 (1991).

    33) A. I. Shnirelman, Usp. Mat. Nauk.30, 265 (1975).

    34) B. V. Chirikov and D. L. Shepelyansky, Phys. Rev. Lett. 74, 518 (1995).

    35) K. M. Frahm and D. L. Shepelyansky, Phys. Rev. Lett. 78, 1440 (1997).

    36) M. C. Gutzwiller, J. Math. Phys.11, 1791 (1970); M. C. Gutzwiller, J. Math. Phys.12,343

    (1971); M. C. Gutzwiller, Chaos in Classical and Quantum Mechanics(Springer-

    17

  • Verlag, Berlin, 1990)

    37) M. L. Mehta and J. des Cloiseaux, Indian J. Pure Appl. Math. 3, 329 (1972).

    38) R. Aurich, A. Backer and F. Steiner, Int. J. Mod. Phys. B 11, 805 (1997).

    39) N. Minami, CRM Proceeding and Lecture Notes 42, 353 (2007).

    40) Seminal article by Rozenzweig and Porter of 1960(see also Ref.3))

    41) F. J. Dyson and M. L. Mehta, J. Math. Phys. 4, 701 (1963).

    42) A. Pandey, Ann. Phys. (N.Y.) 119, 170 (1979).

    43) E. Landau, Archiv der Math. und Physik III 13, 305 (1908).

    44) F. M. de Aguiar, Phys. Rev. E 77,036201 (2008).

    45) H. Makino and S. Tasaki, Prog. Theor. Phys. 5, 929 (2005).

    18

  • FIG.1 Level number variance Σ2(L) of the rectangular billiard systems for (a) 25th, (b) 8th,

    (c) 4th approximations of α = (√

    5 + 1)/2, and (d) α = 1. The solid line denotes the LNV

    of Poisson statistics, Σ2Poisson

    (L) = L. We used energy levels ǭn,m ∈ [4000× 107, 4001× 107].

    The total numbers of levels are (a) 10000457, (b) 10000428, (c)10000266, and (d) 10000162.

    The truncation parameters δ are (a) π × 10−9 (b) π/3× 10−7 and (c) π × 10−9.

    FIG.2 Slope of the level number variance, g(L) = Σ2(L)/L, for (a) 25th, (b) 8th, (c) 4th

    approximations of α = (√

    5 + 1)/2, and (d) α = 1. The solid line, g(L) = 1, corresponds to

    Poisson statistics.

    19

  • FIG.3 Function − ln [1−M(S)] for (a) 25th, (b) 8th, (c) 4th approximations of α = (√

    5 +

    1)/2, and (d) α = 1. The dotted line corresponds to the cumulative Poisson distribution

    M(S) = 1− e−S.

    20

  • FIG.4 Parameter function µ̃(S) for (a) 25th, (b) 8th, (c) 4th approximations of α = (√

    5 +

    1)/2, and (d) α = 1. The solid line, µ̃(S) = 0, corresponds to the Poisson statistics.

    FIG.5 Cumulative NNLSD M(S) for α = 1. We used 999844 levels ǭn,m ∈ [40×106, 41×106]

    for plot (A), 10001872 levels ǭn,m ∈ [4000000× 107, 4000001× 107] for plot (B). We observe

    M(+0) ≃ 0.767 for plot (A) and M(+0) ≃ 0.826 for plot (B). The dotted curve is the

    cumulative Poisson distribution M(S) = 1− e−S

    21

  • FIG.6 Numerical test of Eq.(4.5) for a square billiard (α = 1). The solid line exhibits the

    theoretical prediction, which is valid in the semiclassical limit ǫ → +∞. In each plot, we

    used 107 eigenenergy levels obtained from the unfolded energy range. The error bar exhibits

    the energy range acquiring numerical data.

    22

    Introductionlimiting level number varianceProperties of limiting level number varianceRectangular billiard systemConclusion and Discussion Derivation of Eqs.(2.5ex.3) and (2.5ex.4)